Green Approach for Photocatalytic Cu(II)-EDTA Degradation over TiO2

Jan 15, 2015 - A win–win platform was thus attained, on which Cu was recovered while providing active sites for H2 generation via photoreduction of ...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/est

Green Approach for Photocatalytic Cu(II)-EDTA Degradation over TiO2: Toward Environmental Sustainability Siew Siang Lee,† Hongwei Bai,†,‡ Zhaoyang Liu,§ and Darren Delai Sun*,† †

School of Civil & Environmental Engineering, Nanyang Technological University, Singapore, Singapore 639798 Energy Research Institute@NTU, Nanyang Technological Univeristy, Singapore, Singapore 637141 § Qatar Environment and Energy Research Institute, Qatar Foundation, P.O. Box 5825, Doha, Ad Dawhah, Qatar ‡

S Supporting Information *

ABSTRACT: A green approach was successfully developed to reap three environmental benefits simultaneously: (1) clean water production, (2) hydrogen (H2) generation, and (3) welldispersed in situ Cu2+ recovery for direct TiO2−CuO composite reclamation, by exploiting the synergistic integration of photocatalytic reaction of Cu-EDTA and onedimensional (1D) ultralong and ultrathin TiO2 nanofibers. In this light-initiated system, Cu-EDTA was oxidized by TiO2 thus releasing Cu2+ which was reduced and recovered through uniform adsorption onto the long and porous TiO2 surface. A win−win platform was thus attained, on which Cu was recovered while providing active sites for H2 generation via photoreduction of H2O and enhancing photo-oxidation of remaining intermediate oxidation byproducts. Experimental results showed a H2 generation rate of 251 μmol/h concomitantly with TOC reduction. The used TiO2 nanofibers deposited with Cu were reclaimed directly as the TiO2−CuO composite after a facile heat treatment without additional chemicals and subsequently reusable for photocatalytic treatment of other wastewater (glycerol) to cogenerate H2 and clean water under both UV−visible and visible light. This study expounds a significant advancement through an ingenious integration which enhances the environmental sustainability of Cu-EDTA treatment via TiO2 photocatalysis. It also represents a promising and adoptable approach to synthesize other functional composite nanomaterials in a green manner thus broadening its environmental application spectrum, as it promotes industrial environmental management via waste segregation and motivates research to recover more resources from wastewater.



INTRODUCTION The worldwide burgeoning population, urbanization, and industrialization have resulted in rising concerns among the community today on major environmental issues, in particular (1) water scarcity,1 as well as (2) generation of variable waste or wastewater which are becoming more and more challenging to manage and treat.2 These exigencies have thus obliged more aggressive review, research, and development endeavors in various technologies especially for water/wastewater treatment and reclamation.3−9 Copper(II)- ethylene diamine tetra acetic acid (Cu-EDTA), a type of organometallic wastewater usually generated from textile, electroplating, and nuclear industries, has been one of the major environmental pollutants of concern because it is very stable over a wide range of pH and highly resistant to degradation.10,11 Cu-EDTA wastewater must be treated before discharge, because its persistence in the environment can be detrimental to life as it increases the mobility of Cu elements which can be toxic at high concentration.12 Nonetheless, Cu is also a useful and valuable resource for various applications such as in the electronics and nanomaterials fields.13−15 Unlike the conventional physical-chemical and biological treatments which © XXXX American Chemical Society

entail complex operation and maintenance procedures, high resources demand, as well as challenges in treatment efficiency and post-treatment,16−20 photocatalytic treatment of Cu-EDTA over TiO2 may represent a simpler and more promising technology as (1) it taps into the clean and renewable solar energy for photocatalytic oxidation (PCO) of Cu-EDTA, (2) it spontaneously recovers the Cu2+ via photocatalytic reduction (PCR), (3) it does not require additional chemical, and (4) it does not generate secondary pollutant.11,21,22 Eventually, Cu could be recovered from the used photocatalysts using acid extraction.21 Despite the benefits of concurrent organic degradation and Cu recovery which produce clean water, environmental challenges remain where the post-treatment and complexity of the recovery steps entail intensive use of chemicals, high resources demand and secondary wastewater generation. The challenge is further compounded as exploration to recover the deposited Cu via a greener approach seems Received: October 18, 2014 Revised: January 6, 2015 Accepted: January 15, 2015

A

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

preparation and synthesis procedure as reported in our previous work.13,38,39 Briefly, 7 wt % of PVP and 20 wt % of Ti(oBu)4 were added into a solution of CH3CH2OH and CH3COOH) (V/V = 4:1. Subsequently, the mixture was magnetically mixed for 6 h to obtain a clear sol−gel solution for the electrospinning of TiO2. Facilitated by a syringe pump, the precursor solution was injected at a rate of 0.4 mL/h through a hypodermic syringe with a stainless steel nozzle diameter of 1.1 mm. Under an electrical potential of 1.1 kV/cm, a nonwoven nanofiber webs of PVP/Ti(OBu)4 composite was obtained on a grounded aluminum foil which was located 18 cm away from the nozzle of the syringe. The as-spun nanofibers were calcined in the air at 450 °C for 1 h, at a temperature rising rate of 0.5 °C/min to obtain TiO2 nanofibers. A series of comparative study was conducted to determine the optimum calcinations temperature for TiO2 nanofibers (extensively discussed in the Supporting Information (SI)). Cogeneration of H2 and Clean Water. Concurrent photocatalytic H2 and clean water generation test was carried out in 0.8 mM Cu-EDTA solution containing 0.5 g/L of suspended bare TiO2 nanofibers as the photocatalyst in an inner irradiation type Pyrex reactor with a 400W high pressure Hg lamp (Riko, UVL-400HA) as the UV−visible light source.38 Synthetic Cu-EDTA wastewater solution was prepared by dissolving equal molar amounts of Cu(NO3)2·3H2O and C10H14N2Na2O8·2H2O in DI water. Diluted NaOH and HNO3 were used to adjust the Cu-EDTA solution to pH 4, 6, and 10 to investigate the effect of pH on the reaction. Following a purge with N2 gas for 30 min to deaerate the reactor, dark and photolysis experiments were carried out for 4 h to obtain a baseline for organic removal and H2 generation. In all subsequent photocatalysis reaction tests, the suspension was stirred in the dark for 1 h to achieve a dark adsorption− desorption equilibrium between the organometallic substrates and the photocatalysts prior to light irradiation. An aliquot of 5 mL was sampled and immediately filtered through a Millipore filter of 0.45 μm every 10 min for water quality analysis. The filtered aliquot was measured for its total organic carbon (TOC) and Cu content using a Shimadzu TOC analyzer model TOC-V CSH and inductively coupled plasma optimal emission spectrometry (ICP OES) (Dionex ISC-1000), respectively. A thermal conductivity detector (TCD)-type gas chromatography (Agilent 7890A, HP-PLOT MoleSieve/5A) was used to measure the gas produced from the photocatalytic reaction.29,30 All experiments were carried out in triplicates to get averaged data for analysis. EDTA solution without the presence of Cu ions was tested under the same experimental conditions as comparison. Reclamation of Used TiO2 into TiO2−CuO Composite Nanofibers and Its Usability. The used TiO2 nanofibers with Cu deposited on them (herein after termed as Cu-TiO2 nanofibers (T-C)) were collected and washed briefly with DI water to remove excess organic oxidation byproducts. Subsequently they were dried overnight at 105 °C in an oven and stabilized via calcinations at 450 °C for 1 h to yield TiO2− CuO composite nanofibers (herein after termed as reclaimed TiO2−CuO composite (R-TC)). R-TC were tested for its photocatalytic H2 and clean water production efficiency from glycerol solution (5% v/v) using the same reactor setup for 3 cycles. Photocatalytic efficiency of R-TC under the visible light source (360 W high pressure Na lamp, Riko HNL-360A) was also tested using the same equipment setup.39 The amount of Cu in the reaction solution was determined after each reaction

to be very few. In view of this, one-dimensional (1D) TiO2 nanofibers, which have not been used before for Cu-EDTA photocatalytic treatment, may offer a more sustainable recovery option by facilitating direct reclamation of the used TiO2 photocatalyst with Cu deposited on it, owing to its superior properties: (1) high porosity and long aspect ratio23 which would enhance adsorption and uniform dispersion of reactants and (2) low aggregation tendency which facilitate separation and reusability of the photocatalysts as well as reduce recombination of photogenerated charges.23 Meanwhile, the extent of research on photocatalytic treatment of Cu-EDTA over TiO2 remains limited with other possible sought-after opportunities; one of which is the hydrogen (H2) generation. While EDTA solution itself has been proven to enhance H2 generation,24−27 no study on possible photocatalytic H2 generation from Cu-EDTA was ever reported to date. Past research has shown that Cu was a good cocatalyst and candidate to modify TiO2 thus enhancing its photocatalytic H2 generation significantly.28−31 Hence, in principle, the remaining intermediate oxidation byproducts (after the PCO of CuEDTA) as the hole scavengers32 should enhance the electronsholes separation, thereby enabling further extension of PCR to generate H2 from water (H2O), with a boost of efficiency from the photoreduced Cu on TiO2. In general, derivatives of acetic acid, glyoxylic acid, oxalic acid, glycolic and formic acids were major intermediate degradation products, while NH4+, NO3−, CO2, Cu2+, and some carbon fragments were the common final products of Cu-EDTA photodegradation.32−35 Hypothetically, this may pave an avenue to harvest clean energy simultaneously in addition to clean water production and material reclamation, addressing the looming environmental-energy issues in particular depletion of fossil fuels36,37 and climate change from the use of fossil fuels. Herein for the first time, a green approach was developed to cogenerate H2, clean water, and TiO2−CuO composite from Cu-EDTA wastewater, by exploiting the integration synergy from PCO and PCR extension and the merits of 1D TiO2 nanofibers. Cu-EDTA played a dominant role in promoting the photocatalytic reaction over TiO2 by (1) providing EDTA as the photogenerated holes scavenger thus enhancing the charges separation and (2) providing Cu as the cocatalyst for H2 production, to extend the visible light sensitivity as well as to enhance the separation of the photogenerated charges. What is more, the used TiO2 nanofibers were utilized directly to synthesize TiO2−CuO composite nanomaterials via facile calcinations. Photocatalytic efficiency of the reclaimed composite nanofibers for treatment of other organic pollutant such as glycerol was shown and discussed to predict the application future. In addition, through an in-depth characterization of the reclaimed TiO2−CuO composite, a comprehensive mechanism for this synergistic system was unravelled to bridge the research gap.



EXPERIMENTAL SECTION Materials. Ethanol, acetic acid, tetra n-butyl titanate (Ti(oBu)4), polyvinylpyrolidone (PVP; MW = 1 300 000), copper nitrate (Cu(NO 3 ) 2 ·3H 2 O), sodium-EDTA (C10H14N2Na2O8·2H2O), sodium hydroxide (NaOH), and nitric acid (HNO3) were of analytical grade and used directly without further purification. Deionized (DI) water of 18 μS/cm was used in all reagents and samples preparation. Preparation of TiO2 Nanofibers. TiO2 nanofibers were prepared via electrospinning by adopting the same precursor B

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology cycle using the ICP OES. A leachability test was carried out additionally at pH 2 and pH 12 in diluted HNO3 and diluted NaOH, respectively, to predict the operational pH spectrum for R-TC. A comparison was done with the innovative TiO2−CuO composite nanofibers (6-TC) from our previous work.13 Characterization of Nanofibers (TiO2 and R-TC). The morphologies of photocatalysts were observed and characterized using the field-emission scanning electron microscope (FESEM, JEOL JSM-7600F) coupled with high-end scanning transmission electron microscopy, STEM (working at an accelerated voltage of 30 kV). The elemental content of the photocatalysts were analyzed by an energy dispersive X-ray spectrometer (EDX) (Oxford Instrument, X-Max, 80 mm2) attached to the FESEM (JEOL, JSM-7600F). The microstructure of the nanofibers were analyzed using the high resolution transmission electron microscopy (HRTEM, JEOL JEM-2010) working at an accelerated voltage of 200 kV. The crystal phases were analyzed by the Bruker D8 Advance X-ray diffractometer (XRD) with monochromated high-intensity Cu Kα radiation (λ = 1.5418 Å). The UV−vis spectra were obtained using a Thermo Scientific Evolution 300 UV−vis spectrometer (Thermo Fisher Scientific, Massachusetts, USA). The Brunauer, Emmett, and Teller (BET) specific surface area of the photocatalysts was determined at liquid nitrogen temperature (77 K) using a Micromeritics ASAP 2040 system. The samples were degassed at 200 °C for 3 h prior to BET measurement. The pore size distribution was estimated by employing the Barret-Joyner-Halenda (BJH) method. The Fluorolog-3 spectrofluorometer (Horiba Scientific, New Jersey, USA) was used to conduct the photoluminescence (PL) analysis at an excitation wavelength of 300 nm.40 X-ray photoelectron spectroscopy (XPS) measurements were done using a Kratos Axis Ultra Spectrometer with a monochromic Al Kα source at 1486.7 eV, a voltage of 15 kV, and an emission current of 10 mA; where the carbonaceous C 1s line at 284.6 eV was used as the reference to calibrate the binding energies.

Figure 1. (a) Photocatalytic TOC removal (%) and H2 generation rate for Cu-EDTA and EDTA and (b) photocatalytic removal of TOC (%), Cu (%), and H2 generation over 4 h reaction in Cu-EDTA.

literature, however, with no further explanation included.11 The effect of pH was also investigated, and the results are depicted in Table 1. The shortest time was required at pH 4 for Table 1. Effect of Solution pH on Photocatalytic Reaction



RESULTS AND DISCUSSION Cogeneration of H2 and Clean Water. H2 and clean water generation from Cu-EDTA was investigated using electrospun TiO2 nanofibers under UV−visible irradiation. Its efficiency is compared against that of a pure EDTA solution without the presence of Cu2+ ions to gain more insights of the reaction. TOC and Cu2+ removal were monitoring parameters for the degradation of Cu-EDTA thus the clean water generation. In the case of EDTA, only TOC was monitored. As shown in Figure 1(a), at the end of the 4 h irradiation, nearly 2 times more H2 generation rate of approximately 251 μmol/h was obtained from Cu-EDTA than that from EDTA over TiO2 nanofibers. TOC removal from Cu-EDTA was observed to be nearly 1.5 times higher than that from EDTA. No appreciable H2 generation and TOC removal was observed both in the dark and during photolysis (in the absence of photocatalysts) signifying the essential role of TiO2 nanofibers for photocatalytic treatment of both Cu-EDTA and EDTA (the efficiency of TiO2 nanofibers in EDTA was discussed in the SI). Figure 1(b) shows the continuous percentage removal of TOC and Cu2+ from Cu-EDTA over TiO2 nanofibers throughout the 4 h reaction. The more superior H2 generation rate and TOC removal in Cu-EDTA than EDTA have signified the dominant role of Cu from Cu-EDTA in promoting the photocatalytic reaction. Similar observation on the positive effect of Cu2+ on photodegradation of EDTA was reported before in the

pH condition (Cu-EDTA)

time to achieve 90% Cu reduction (min)

TOC removal after 4 h reaction (%)

H2 evolution (μmol/h/g)

pH 4 pH 6 pH 10

10 40 50

57.5 71.0 70.0

2293.0 1930.3 2196.6

90% of the Cu2+ to be removed from the solution. The protonated TiO2 surface under acidic conditions facilitated the adsorption of the anionic Cu-EDTA,17,32 thus promoting the PCO of the adsorbed Cu-EDTA by the photogenerated holes and subsequently the decomplexation and PCR of Cu2+ on TiO2 by the photogenerated electrons. Meanwhile, possible protonation of Cu-EDTA could happen at low pH, leading to decomplexation in the bulk solution prior to adsorption on TiO2.41 TOC removal, nevertheless, was more favorable at pH 6 and 10 compared to that at pH 4. Even though the presence of OH− would pose a competition with Cu-EDTA for adsorption and reaction sites on TiO2,17 its formation as OH radical (OH•) by the photogenerated holes19 would eventually facilitate Cu-EDTA oxidation in the bulk solution. The low TOC removal at pH 4 further suggested that the overall TOC reduction could be enhanced by OH−, in addition to the adsorption and oxidation by the holes. In the case where PCO of Cu-EDTA took place in the bulk solution, the PCR of Cu2+ could be delayed due to the additional path required for the decomplexed Cu2+ to reach and adsorb onto TiO2, as evidenced by the longer Cu2+ removal at pH 6 and 10. However, negatively charged TiO2 under basic conditions might counter this effect by favoring the adsorption of Cu2+.17,42 Overall, regardless of solution pH, approximately C

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

photocatalytic reaction.49,50 The reduced BET specific surface area in R-TC compared to TiO2 nanofibers (Table S1) is consistent with the observation in Figure S5, further affirming the deposition of Cu elements on and within TiO2 nanofibers. Meanwhile, the increased BET specific surface area of R-TC compared to T-C has suggested that the organics have been removed during calcinations. The HRTEM image (Figure S6(a)) further elucidates the closely contacted TiO2−CuO heterojunctions. Both the (101) plane of anatase TiO2 and the (111) plane of CuO were well-evidenced by the lattice fringes of 0.325 and 0.253 nm, respectively.40,51,52 The XRD pattern (inset of Figure S6(a)) revealed that in addition to anatase TiO2 (JCPDS file no. 21-1272),40 CuO was present. A close-up XRD pattern plotted across the region of 2θ = 30°−45° (Figure S6(b)) provided a clear view where the peaks observed at 2θ = 36.5° and 2θ = 38.8° could be ascribed to the (111) plane and (002) plane of CuO, respectively,52 inconsistent with the observed HRTEM image. The absorbance spectra of TiO2 and R-TC in the UV−vis light region of between 350 and 800 nm (Figure S7) illustrated that the absorption edge in the UV light region has been red-shifted for R-TC, suggesting the successful incorporation of the CuO into TiO2 nanofibers as wellwitnessed in the HRTEM image and XRD patterns in Figure S6. In addition, the photoresponse in the visible light region was significantly intensified likely due to the presence of CuO which has a small bandgap of 1.2 eV thus capable of absorbing both the UV and visible regions of the solar spectrum.53 The enhanced sensitivity and absorption in the visible light region could enhance the photogeneration of electrons and holes for photocatalytic reaction. The addition of CuO has resulted in a lower PL spectra by R-TC than that by TiO2 nanofibers (Figure S3). A lower PL spectra intensity implies a lower recombination tendency of the photogenerated charges.40 The energy potential differences which existed between the conduction band (CB) of TiO2 and that of the CuO have promoted the migration of the photogenerated electrons from the CB of TiO2 to that of CuO. This has further corroborated the advantage of good elemental dispersion and close proximity between TiO2 and CuO heterojunctions as shown in Figure S5 and Figure S6; which implied that the separation and hence the life span and utilization of photogenerated charges could be enhanced for photocatalytic reaction. The chemical composition and elemental states of R-TC were analyzed using the XPS analysis (Figure S8). The binding energy for the C 1s peak at 284.6 eV was used as the reference for calibration. The survey spectrum (Figure S8(a)) witnesses the existence of Ti, Cu, and O in the regenerated TiO2 nanofibers. The characteristic peaks for Ti4+ at the binding energy of 458.0 and 464.0 eV (Figure S8(b)) were indicative of Ti 2p3/2 and Ti 2p1/2, respectively. The peaks at the binding energy of 529.5 eV which was evidence of O 1s in TiO2 and CuO29 (Figure S8(c)) implied contamination by atmospheric hydroxides onto the surface of the composite nanofibers.29 The valence state of Cu was +2, as corroborated by the peaks around the binding energy of 933.9 and 953.9 eV ascribing to the Cu 2p3/2 and Cu 2p1/2 of Cu2+ in CuO,31 respectively (Figure S8(d)). The existence of Cu2+ from CuO in the composite nanofibers was further confirmed by the characteristic shakeup satellite peaks around 942.0 and 962.0 eV.54 On the basis of the characterizations and analysis, it is reasonable to state that Cu has been successfully loaded into the TiO2 nanofibers and, after calcinations, formed TiO2−CuO composite (R-TC) which exhibited rather similar properties as our previously reported innovative TiO2−CuO nanofibers

50% TOC and 90% of Cu2+ removal was achieved within 1 h. Meanwhile, the combined effect of H+ as the precursor to H2 production30,43 as well as photoreduction of Cu2+ as reduction sites on TiO244 seemed to have promoted H2 generation at pH 4; despite the small differences on H2 evolution rate between all pH values. This part of the work showed that H2 and clean water can be generated concurrently and more efficiently from Cu-EDTA through PCO of the remaining oxidation byproducts and PCR of H2O or H+. Reclamation of T-C and Its Characterization. At the end of the photocatalytic reaction, T-C were filtered from the reacted solution for direct synthesis into R-TC via facile calcinations at 450 °C for 1 h, to regenerate the photocatalysts by removing any excess organics byproduct and to oxidize Cu into CuO. In-depth characterizations of TiO2 and R-TC were carried out following the photocatalytic reaction in the CuEDTA solution to compare, to gain an insight into any physicochemical properties changes, and to verify the formation of R-TC. Figure 2(a) shows that TiO2 nanofibers possessed a

Figure 2. (a) High magnification and low magnification (inset) FESEM images of ultralong and ultrathin electrospun 1D TiO2 nanofibers and (b) high magnification and low magnification (inset) FESEM images of reclaimed TiO2−CuO composite nanofibers (RTC).

rough and porous structure given by the loss of PVP during calcinations,45 as well as a high aspect ratio of around 30−50 (inset). This would render the photocatalysts with high specific surface area thus reaction sites for adsorption and photocatalytic reaction of the reactants.46 As shown in Figure 2(b), the sustained length and diameter in R-TC will be advantageous for the subsequent application and recovery because these properties facilitate interparticle charges transfer and prevent agglomeration of the nanofibers.47 A porous structure analysis using the N2 adsorption−desorption isotherm technique showed that R-TC has remained a mesoporous property like TiO2 nanofibers, as characterized by the type IV isotherm curve with obvious H3-type hysteresis (Figure S4).48 The mesoporous nature of R-TC was further corroborated by the BJH analysis where a narrow pore size distribution of between 2 and 20 nm was observed (inset of Figure S4).48 Figure S5(a) shows a high-magnification STEM-FESEM image of a single strain of R-TC. Figure S5(b) depicts the possible aggregation of Cu which was however generally rare as seen in the inset of Figure S5(b), while Figure S5(c) confirms the presence of both Ti and Cu on R-TC. The mapping images as shown in Figures S5(d)-(f) further supported the coexistence of well-dispersed Ti, Cu, and O elements, facilitated by the mesoporosity and the ultralong 1D structure of the nanofibers (Figure 2). A good Cu dispersion is highly desirable to attain an optimized elemental contact between TiO2 and CuO heterojunctions which thus would enhance the transfer and separation of the photogenerated charges for a more efficient D

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

Figure 3. (a) Photocatalytic H2 generation from 5%(v/v) glycerol using R-TC under UV−visible and visible lights and (b) comparison of H2 evolution rate from 5% v/v glycerol under UV−visible and visible lights between R-TC and innovative TiO2−CuO composite (6-TC), and (c) reusability experiment for photocatalytic H2 generation by R-TC.

(6TC).13 Hence, R-TC could be a potential nanomaterial candidate for a photocatalytic reaction. Unlike the conventionally adopted method such as wet-impregnation, hydrothermal, and sol−gel which entail use of additional chemical or complex synthesis steps to incorporate Cu on the surface of the TiO2 photocatalysts,29,30,49,55 the recovered Cu from Cu-EDTA wastewater can be utilized directly as the precursor to synthesize TiO2−CuO composite almost instantly in a more sustainable manner by tapping into the benefits of 1D TiO2 nanofibers. With this, not only the nonrenewable chemical resource could be conserved, there was also opportunity for attaining zero waste discharge from the wastewater treatment, assuming that a proper drain segregation has been implemented upstream at the point of wastewater generation to prevent mixing of a variety of waste components. R-TC - Photocatalytic Efficiency and Stability. The RTC were tested for their photocatalytic function to cogenerate H2 and clean water generation from glycerol, a primary waste product from the biodiesel production industry.28 The mechanism involved in the photocatalytic treatment of glycerol using TiO2 to cogenerate H2 and clean water has been proposed elsewhere.28,56 Despite the reduced specific surface area, R-TC exhibited efficient H 2 generation rate of approximately 2627 μmol/h under UV−visible light and 407 μmol/h under visible light, as shown in Figure 3(a). This is almost comparable with the capability of 6TC which showed 2647 μmol/h and 562 μmol/h H2 evolution rate from glycerol under UV−visible and visible light, respectively (Figure 3(b)). Rhoads K. et al. (2004) however reported that the used TiO2 exhibited similar efficiency with or without adsorbed Cu, thus implying the significant use of 1D TiO2 nanofibers to facilitate the green synthesis of usable R-TC directly.21 It is thus postulated that the synergy between all the enhanced physicochemical properties of R-TC could have driven the net photocatalytic reaction.39 Albeit offering a similar role as

Cu0 in T-C before calcinations, CuO in R-TC exhibited enhanced charge separation and stability, thus enhancing photocatalytic H2 production.13,30,44,57−59 Advantage of R-TC over T-C is elaborated in the SI. Repeated use of R-TC was feasible owing to its sustainable 1D long nanofibrous structure which thus have enabled the ease of recovery and reuse, corroborated by the FESEM-STEM images (Figure S9) showing sustained morphology of the used R-TC. XRD and EDX analysis for used R-TC are available in the SI. Negligible reduction in photocatalytic H2 generation was observed in the subsequent two cycles (Figure 3(c)). The specific surface area could be easily recovered via facile calcinations at 450 °C overnight (Table S1). Moreover, Cu2+ ion was constantly below the detection limit of 1.0 mg/L in the reacting solution after every cycle (Table S6), complying with the industrial wastewater effluent standard for Cu as specified by the US EPA.60 While the leachability study reveals that more than 1.0 mg/L Cu2+ could be leached out at pH 2, further study is undergoing to identify the optimization possibility in order to extend the operational pH range for R-TC as well as to build the fundamental of this approach. In essence, the exhibited photocatalytic efficiency and stability imply the potential of this green approach to synthesize TiO2−CuO composite, which subsequently could promote industrial symbiosis when the reclaimed photocatalyst is used for remediation of wastewater from another industry. Mechanism Explanation. Based on the experimental data and characterization studies, the mechanism of producing clean water, H2, and R-TC simultaneously via the photocatalytic treatment of Cu-EDTA was postulated and illustrated in Scheme 1. The mechanism comprised the reaction that takes place in the bulk solution and on the photocatalysts. Upon harvesting of the photon energy from the simulated solar irradiation, electrons and holes responsible for the photocatalytic reaction were generated in the CB and valence band E

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology

generation rate from Cu-EDTA was more efficient than that from EDTA. While the Cu0 could also be possibly oxidized by the holes into Cu2+,19,44 and while the PCR of the Cu2+ into Cu0 would result in the competition for electrons to generate H2,44 the overall net effect of Cu0/Cu2+ and Cu2+/Cu0 redox reactions were seen to be insignificant and would become negligible at one stage, especially as the reaction progressed beyond 1 h after which almost 90% of Cu2+ were removed along with a continuous H2 evolution. Briefly, this study has shown that clean water, clean energy fuel, and R-TC could be cogenerated from photocatalytic treatment of Cu-EDTA through the excellent synergy rendered by extending PCO and PCR of Cu-EDTA coupled with TiO2 nanofibers of superior physicochemical properties. With the elimination of (1) prior complex synthesis process to modify and enhance TiO2, (2) sacrificial reagent to enhance reaction efficiency for clean water and clean energy production, and (3) additional chemical to recover Cu, this approach has enhanced the sustainability of Cu-EDTA treatment via TiO2 photocatalysis significantly. What is more, this ingenious synergistic integration represents an adoptable approach to synthesize other functional composite nanomaterials from other organometallic wastewater streams in a green manner, possibly broadening its environmental application prospect. While this approach is being pursued; the monitoring of and identification of suitable treatment technology for NH4+, NO3−, and dissolved organic carbon may be necessary to close the loop in the later stage as this research matures, since their excessive existence in the water body would adversely impact the health of aquatic ecosystem and its surrounding living community.

Scheme 1. Mechanism behind the Cogeneration of H2, Clean Water, and R-TC from the Photocatalytic Treatment of CuEDTA Using 1D TiO2 Nanofibers

(VB) of the TiO2 nanofibers, respectively. The photogenerated holes would react with the hydroxyl groups in the aqueous phase or on the surface of the photocatalysts, yielding the reactive oxidation species (ROS) which possess strong oxidation ability. The PCO of Cu-EDTA was initiated either by the ROS in the bulk solution, or by the photogenerated holes, and/or the ROS on the surface of the TiO2 nanofibers where Cu-EDTA was adsorbed onto. Subsequently, decomplexation of the Cu2+ from Cu-EDTA complexes would take place, followed by the adsorption of Cu 2+ onto TiO 2 nanofibers.10,32 Cu2+ ions in the bulk solution however would need to travel in order to access the photocatalysts prior to adsorption, especially under high pH condition where adsorption of Cu-EDTA was challenged. Thereafter, Cu2+ could be reduced into Cu metal (Cu0) by the photogenerated electrons to yield T-C.42,44 XRD pattern and leachability analysis for T-C are available in the SI. A uniform dispersion and adsorption of Cu2+ was facilitated by the mesoporosity and long nanofibrous structure of the TiO2 nanofibers.42 H2 could be harvested by extending PCO of the remaining intermediate oxidation byproducts and PCR of H2O or H+. The extensive analysis of reaction products from the PCO of Cu-EDTA which was well-discussed in other literature is not included in this research.32,33,35,61 Meanwhile, since the redox potential of Cu0 is more electronegative than H2O reduction potential, the migration of photogenerated electrons was promoted from the CB of TiO2 to Cu0 following which electrons would be rapidly and efficiently used up for H2O reduction into H2.44 Under acidic conditions, additional H+ would enhance the H2 generation. On the surface of the TiO2 nanofibers, Cu0 rendered several advantages as follows: (1) it served as an efficient reduction cocatalyst for H2 production, (2) it promoted the separation of photogenerated charges thus a longer lifetime of the charge carriers, and (3) it extended the light sensitivity and absorption thus the charges photogeneration ability of TiO2 nanofibers.44 Hence, the H 2



ASSOCIATED CONTENT

S Supporting Information *

Characterization and verification studies as noted in the text available: Figures S1−S10 and Tables S1−S6. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Phone: 65-6790-6273. Fax: 65-6790-0676. E-mail: ddsun@ ntu.edu.sg. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank the Clean Energy Research Programme under National Research Foundation of Singapore for their research grant (Grant No. NRF2007EWT-CERP01-0420) support for this work and the support received from Public Utilities Board (PUB) of Singapore. The Scholarship provided by NTU is appreciated.



REFERENCES

(1) Sophocleous, M. Global and regional water availability and demand: Prospects for the future. Nat. Resour. Res. 2004, 13 (2), 61− 75. (2) Thavasi, V.; Singh, G.; Ramakrishna, S. Electrospun nanofibers in energy and environmental applications. Energy Environ. Sci. 2008, 1 (2), 205−221. (3) Adav, S. S.; Lee, D. J.; Lai, J. Y. Treating chemical industries influent using aerobic granular sludge: Recent development. J. Taiwan Inst. Chem. Eng. 2009, 40 (3), 333−336.

F

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology (4) Bora, T.; Dutta, J. Applications of nanotechnology in wastewater treatment-A review. J. Nanosci. Nanotechnol. 2014, 14 (1), 613−626. (5) Li, W. W.; Yu, H. Q.; He, Z. Towards sustainable wastewater treatment by using microbial fuel cells-centered technologies. Energy Environ. Sci. 2014, 7 (3), 911−924. (6) Petrović, M.; Gonzalez, S.; Barceló, D. Analysis and removal of emerging contaminants in wastewater and drinking water. TrAC, Trends Anal. Chem. 2003, 22 (10), 685−696. (7) Qu, X.; Alvarez, P. J. J.; Li, Q. Applications of nanotechnology in water and wastewater treatment. Water Res. 2013, 47 (12), 3931− 3946. (8) Theron, J.; Walker, J. A.; Cloete, T. E. Nanotechnology and water treatment: Applications and emerging opportunities. Crit. Rev. Microbiol. 2008, 34 (1), 43−69. (9) Yang, B. M.; Huang, C. J.; Lai, W. L.; Chang, C. C.; Kao, C. M. Development of a three-stage system for the treatment and reclamation of wastewater containing nano-scale particles. Desalination 2012, 284, 182−190. (10) Cho, I. H.; Shin, I. S.; Yang, J. K.; Lee, S. M.; Shin, W. T. Removal of Cu(II)-EDTA complex using TiO2/solar light: The effect of operational parameters and feasibility of solar light application. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2006, 41 (6), 1027−1041. (11) Park, E.-H.; Jung, J.; Chung, H.-H. Simultaneous oxidation of EDTA and reduction of metal ions in mixed Cu(II)/Fe(III)−EDTA system by TiO2 photocatalysis. Chemosphere 2006, 64 (3), 432−436. (12) Nybroe, O.; Brandt, K. K.; Ibrahim, Y. M.; Tom-Petersen, A.; Holm, P. E. Differential bioavailability of copper complexes to bioluminescent Pseudomonas fluorescens reporter strains. Environ. Toxicol. Chem. 2008, 27 (11), 2246−2252. (13) Lee, S. S.; Bai, H.; Liu, Z.; Sun, D. D. Novel-structured electrospun TiO2/CuO composite nanofibers for high efficient photocatalytic cogeneration of clean water and energy from dye wastewater. Water Res. 2013, 47 (12), 4059−4073. (14) Braun, A. E. Electroplating. Semicond. Int. 2003, 26 (8), 160. (15) Taberna, P. L.; Mitra, S.; Poizot, P.; Simon, P.; Tarascon, J. M. High rate capabilities Fe3O4-based Cu nano-architectured electrodes for lithium-ion battery applications. Nat. Mater. 2006, 5 (7), 567−573. (16) Chang, L. Y. A waste minimization study of a chelated copper complex in wastewater - Treatability and process analysis. Waste Manage. 1995, 15 (3), 209−220. (17) Chung, H. H.; Rho, J. S. Photooxidation of EDTA on TiO2. J. Ind. Eng. Chem. 1999, 5 (4), 261−267. (18) Chaudhary, A. J.; Donaldson, J. D.; Grimes, S. M.; Ul-Hassan, M.; Spencer, R. J. Simultaneous recovery of heavy metals and degradation of organic species - Copper and ethylenediaminetetraacetic acid (EDTA). J. Chem. Technol. Biotechnol. 2000, 75 (5), 353− 358. (19) Kim, S. J.; Lee, H. G.; Lee, J. K.; Lee, E. G. Photoredox properties of ultrafine rutile TiO2 acicular powder in aqueous 4chlorophenol, Cu-EDTA and Pb-EDTA solutions. Appl. Catal., A 2003, 242 (1), 89−99. (20) Wu, L.; Wang, H.; Lan, H.; Liu, H.; Qu, J. Adsorption of Cu(II)−EDTA chelates on tri-ammonium-functionalized mesoporous silica from aqueous solution. Sep. Purif. Technol. 2013, 117 (0), 118− 123. (21) Rhoads, K. R.; Davis, A. P. Metal recovery and catalyst reuse from the photocatalytic oxidation of copper-ethylenediaminetetraacetic acid. J. Environ. Eng. 2004, 130 (4), 425−431. (22) Kabra, K.; Chaudhary, R.; Sawhney, R. L. Treatment of hazardous organic and inorganic compounds through aqueous-phase photocatalysis: A review. Ind. Eng. Chem. Res. 2004, 43 (24), 7683− 7696. (23) Choi, S. K.; Kim, S.; Lim, S. K.; Park, H. Photocatalytic comparison of TiO2 nanoparticles and electrospun TiO2 nanofibers: Effects of mesoporosity and interparticle charge transfer. J. Phys. Chem. C 2010, 114 (39), 16475−16480.

(24) Zhou, H.; Qu, Y.; Zeid, T.; Duan, X. Towards highly efficient photocatalysts using semiconductor nanoarchitectures. Energy Environ. Sci. 2012, 5 (5), 6732−6743. (25) Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renewable Sustainable Energy Rev, 2007, 11 (3), 401−425. (26) Sabaté, J.; Cervera-March, S.; Simarro, R.; Giménez, J. A comparative study of semiconductor photocatalysts for hydrogen production by visible light using different sacrificial substrates in aqueous media. Int. J. Hydrogen Energy 1990, 15 (2), 115−124. (27) Kim, G.; Choi, W. Charge-transfer surface complex of EDTATiO2 and its effect on photocatalysis under visible light. Appl. Catal., B 2010, 100 (1−2), 77−83. (28) Yu, J.; Ran, J. Facile preparation and enhanced photocatalytic H2-production activity of Cu(OH) 2 cluster modified TiO2. Energy Environ. Sci. 2011, 4 (4), 1364−1371. (29) Xu, S.; Du, A. J.; Liu, J.; Ng, J.; Sun, D. D. Highly efficient CuO incorporated TiO2 nanotube photocatalyst for hydrogen production from water. Int. J. Hydrogen Energy 2011, 36 (11), 6538−6545. (30) Xu, S.; Sun, D. D. Significant improvement of photocatalytic hydrogen generation rate over TiO2 with deposited CuO. Int. J. Hydrogen Energy 2009, 34 (15), 6096−6104. (31) Wu, N. L.; Lee, M. S. Enhanced TiO2 photocatalysis by Cu in hydrogen production from aqueous methanol solution. Int. J. Hydrogen Energy 2004, 29 (15), 1601−1605. (32) Yang, J. K.; Davis, A. P. Photocatalytic oxidation of Cu(II) EDTA with illuminated TiO2: Mechanisms. Environ. Sci. Technol. 2000, 34 (17), 3796−3801. (33) Zhao, X.; Guo, L.; Zhang, B.; Liu, H.; Qu, J. Photoelectrocatalytic oxidation of CuII-EDTA at the TiO2 electrode and simultaneous recovery of CuII by electrodeposition. Environ. Sci. Technol. 2013, 47 (9), 4480−4488. (34) Krapfenbauer, K.; Getoff, N. Comparative studies of photo- and radiation-induced degradation of aqueous EDTA. Synergistic effects of oxygen, ozone and TiO2 (acronym: CoPhoRaDe/EDTA). Radiat. Phys. Chem. 1999, 55 (4), 385−393. (35) Sillanpäa,̈ M. E. T.; Kurniawan, T. A.; Lo, W. H. Degradation of chelating agents in aqueous solution using advanced oxidation process (AOP). Chemosphere 2011, 83 (11), 1443−1460. (36) Schloegl, R. Energy: Fuel for thought. Nat. Mater. 2008, 7 (10), 772−774. (37) Chen, X.; Shen, S.; Guo, L.; Mao, S. S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110 (11), 6503−6570. (38) Lee, S. S.; Bai, H.; Liu, Z.; Sun, D. D. Electrospun TiO2/SnO2 nanofibers with innovative structure and chemical properties for highly efficient photocatalytic H2 generation. Int. J. Hydrogen Energy 2012, 37 (14), 10575−10584. (39) Lee, S. S.; Bai, H.; Liu, Z.; Sun, D. D. Optimization and an insightful properties-Activity study of electrospun TiO2/CuO composite nanofibers for efficient photocatalytic H2 generation. Appl. Catal., B 2013, 140−141, 68−81. (40) Ng, J.; Xu, S.; Zhang, X.; Yang, H. Y.; Sun, D. D. Hybridized nanowires and cubes: A novel architecture of a heterojunctioned TiO2/SrTiO3 thin film for efficient water splitting. Adv. Funct. Mater. 2010, 20 (24), 4287−4294. (41) Yang, J. K.; Davis, A. P. Competitive adsorption of Cu(II)EDTA and Cd(II)-EDTA onto TiO2. J. Colloid Interface Sci. 1999, 216 (1), 77−85. (42) Yang, J. K.; Davis, A. P. Photocatalytic oxidation of Cu(II)EDTA with illuminated TiO2: Kinetics. Environ. Sci. Technol. 2000, 34 (17), 3789−3795. (43) Choi, W. Photocatalytic hydrogen production using surface modified titania nanoparticles. Proc. SPIE 2007, DOI: 10.1117/ 12.737292. (44) Xu, S.; Ng, J.; Du, A. J.; Liu, J.; Sun, D. D. Highly efficient TiO2 nanotube photocatalyst for simultaneous hydrogen production and G

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Article

Environmental Science & Technology copper removal from water. Int. J. Hydrogen Energy 2011, 36 (11), 6538−6545. (45) Li, D.; McCann, J. T.; Xia, Y.; Marquez, M. Electrospinning: A simple and versatile technique for producing ceramic nanofibers and nanotubes. J. Am. Ceram. Soc. 2006, 89 (6), 1861−1869. (46) Pan, J. H.; Dou, H.; Xiong, Z.; Xu, C.; Ma, J.; Zhao, X. S. Porous photocatalysts for advanced water purifications. J. Mater. Chem. 2010, 20 (22), 4512−4528. (47) Pan, J. H.; Sun, D. D.; Lee, C.; Kim, Y. J.; Lee, W. I. Effect of calcination temperature on the textural properties and photocatalytic activities of highly ordered cubic mesoporous WO3/TiO2 films. J. Nanosci. Nanotechnol. 2010, 10 (7), 4747−4751. (48) Sing, K. S. W. Physisorption of nitrogen by porous materials. J. Porous Mater. 1995, 2 (1), 5−8. (49) Yoong, L. S.; Chong, F. K.; Dutta, B. K. Development of copperdoped TiO2 photocatalyst for hydrogen production under visible light. Energy 2009, 34 (10), 1652−1661. (50) Bokhimi, X.; Morales, A.; Novaro, O.; López, T.; Chimal, O.; Asomoza, M.; Gómez, R. Effect of Copper Precursor on the Stabilization of Titania Phases, and the Optical Properties of Cu/ TiO2 Prepared with the Sol-Gel Technique. Chem. Mater. 1997, 9 (11), 2616−2620. (51) Gouma, P. I.; Lee, J. Tailored 3D CuO nanogrid formation. J. Nanomater. 2011, 2011. (52) Mor, G. K.; Varghese, O. K.; Wilke, R. H. T.; Sharma, S.; Shankar, K.; Latempa, T. J.; Choi, K. S.; Grimes, C. A. p-type Cu-Ti-O nanotube arrays and their use in self-biased heterojunction photoelectrochemical diodes for hydrogen generation. Nano Lett. 2008, 8 (7), 1906−1911. (53) Yu, H.; Yu, J.; Liu, S.; Mann, S. Template-free hydrothermal synthesis of CuO/Cu2O composite hollow microspheres. Chem. Mater. 2007, 19 (17), 4327−4334. (54) Li, G.; Dimitrijevic, N. M.; Chen, L.; Rajh, T.; Gray, K. A. Role of surface/interfacial Cu2+ sites in the photocatalytic activity of coupled CuO-TiO2 nanocomposites. J. Phys. Chem. C 2008, 112 (48), 19040−19044. (55) Kim, K. H.; Ihm, S. K. Characteristics of titania supported copper oxide catalysts for wet air oxidation of phenol. J. Hazard. Mater. 2007, 146 (3), 610−616. (56) Lalitha, K.; Sadanandam, G.; Kumari, V. D.; Subrahmanyam, M.; Sreedhar, B.; Hebalkar, N. Y. Highly stabilized and finely dispersed Cu2O/TiO2: A promising visible sensitive photocatalyst for continuous production of hydrogen from glycerol:water mixtures. J. Phys. Chem. C 2010, 114 (50), 22181−22189. (57) Xu, S.; Ng, J.; Zhang, X.; Bai, H.; Sun, D. D. Fabrication and comparison of highly efficient Cu incorporated TiO 2 photocatalyst for hydrogen generation from water. Int. J. Hydrogen Energy 2010, 35 (11), 5254−5261. (58) Bandara, J.; Udawatta, C. P. K.; Rajapakse, C. S. K. Highly stable CuO incorporated TiO2 catalyst for photocatalytic hydrogen production from H2O. Photochem. Photobiol. Sci. 2005, 4 (11), 857− 861. (59) Helaïli, N.; Bessekhouad, Y.; Bouguelia, A.; Trari, M. Visible light degradation of Orange II using xCuyOz/TiO2 heterojunctions. J. Hazard. Mater. 2009, 168 (1), 484−492. (60) Chen, R. H.; Chai, L. Y.; Wang, Y. Y.; Liu, H.; Shu, Y. D.; Zhao, J. Degradation of organic wastewater containing Cu-EDTA by Fe-C micro-electrolysis. Trans. Nonferrous Met. Soc. China 2012, 22 (4), 983−990. (61) Madden, T. H.; Datye, A. K.; Fulton, M.; Prairie, M. R.; Majumdar, S. A.; Stange, B. M. Oxidation of metal-EDTA complexes by TiO2 photocatalysis. Environ. Sci. Technol. 1997, 31 (12), 3475− 3481.

H

DOI: 10.1021/es504711e Environ. Sci. Technol. XXXX, XXX, XXX−XXX